10.04.2014 Views

Nematic liquid crystal waveguide arrays

Nematic liquid crystal waveguide arrays

Nematic liquid crystal waveguide arrays

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

OPTO-ELECTRONICS REVIEW 13(2), 107–112<br />

7 th International Workshop on Nonlinear Optics Applications<br />

<strong>Nematic</strong> <strong>liquid</strong> <strong>crystal</strong> <strong>waveguide</strong> <strong>arrays</strong><br />

K.A. BRZD¥KIEWICZ *1 , M.A. KARPIERZ 1 , A. FRATALOCCHI 2 , G. ASSANTO 2 ,<br />

and E. NOWINOWSKI-KRUSZELNICKI 3<br />

1 Faculty of Physics, Warsaw University of Technology, 75 Koszykowa Str., 00–662 Warsaw, Poland<br />

2 NooEL-Nonlinear Optics and Optoelectronics Laboratory, University “Roma Tre”,<br />

Via della Vasca Navale 84, 00–146 Rome, Italy<br />

3 Institute of Applied Physics, Military University of Technology, 2 Kaliskiego Str., 00–908 Warsaw, Poland<br />

We investigate linear and nonlinear light propagation in a voltage-tunable array of <strong>waveguide</strong> channels in undoped nematic<br />

<strong>liquid</strong> <strong>crystal</strong>s. This novel geometry, based on a photonic structure with a periodic modulation of refractive index controlled<br />

by an electric field, offers a wealth of possibilities for the study of discrete optical phenomena. The structure, in conjunction<br />

with a giant, non-resonant and voltage-dependent reorientational nonlinearity, allows us to drive the system from bulk diffraction<br />

to discrete propagation. Theoretical and experimental investigations, carried out with near infrared light wavelength<br />

and powers of a few milliWatts, show the possibility of transverse light localization, resulting in discrete spatial solitons.<br />

Such array, with its voltage- and light-adjustable guided-wave confinement and coupling, exhibits potentials for the realization<br />

of multifunctional routers and all-optical signal processors with nematic <strong>liquid</strong> <strong>crystal</strong>s.<br />

Keywords: nonlinear optics, solitons, self-action effects, optical nonlinearity in <strong>liquid</strong> <strong>crystal</strong>s, discrete diffraction.<br />

1. Introduction<br />

Linear and nonlinear effects in discrete systems such as<br />

photonic structures with a periodic modulation of the refractive<br />

index have became the subject of intensive investigations.<br />

Particular attention has been devoted to the generation<br />

of discrete solitons [1,2] which are excellent candidates<br />

for novel all-optical switching devices and networks<br />

[3–6]. Similarly to continuous systems, stable spatial<br />

solitons can be formed in discrete structures when self focusing<br />

is strong enough to balance discrete diffraction.<br />

Moreover, discrete solitons can be viewed as due to a nonlinear<br />

response which detunes a <strong>waveguide</strong> from the neighbouring<br />

ones. General properties of discrete optical solitons<br />

were first described by Christodoulides et al. referring to a<br />

cubic Kerr nonlinearity [7] and later studied with reference<br />

to other nonlinear materials, including photorefractive and<br />

parametric <strong>crystal</strong>s [1,8]. To the date, they were observed<br />

experimentally in AlGaAs [9–12], in silica [13,14], in<br />

SBN-<strong>crystal</strong>s [15,16], in Lithium Niobate [17], and recently<br />

also in nematic <strong>liquid</strong> <strong>crystal</strong>s (NLC) [18]. The latter<br />

were proven to be very promising for nonlinear optics<br />

[19–21]. Their inexpensive and consolidated technology,<br />

huge birefringence, giant reorientational nonlinearity and<br />

low power requirements, have allowed us demonstrating<br />

that spatial solitons can be effectively generated at mW<br />

power levels, in bulk as well as in planar <strong>waveguide</strong> configurations,<br />

with propagation distances of a few millimeters<br />

[20]. Moreover, the large electro-optic response of NLC allows<br />

for creation of flexible and voltage-tunable architectures,<br />

useful for the study of discrete phenomena. In this<br />

communication, we illustrate a one-dimensional periodic<br />

photonic structure with refractive index contrast controlled<br />

by an external electric field. We present experimental and<br />

theoretical analyses of both linear and nonlinear discrete<br />

propagation of infrared-light in <strong>arrays</strong> of identical<br />

weakly-coupled <strong>waveguide</strong>s in undoped nematic <strong>crystal</strong>s,<br />

paying particular attention to the choice of geometric parameters.<br />

A sketch of the device is drawn in Fig. 1. A few-micron<br />

thick layer of PCB (4-cyano-4'-n-pentylbiphenyl) nematic<br />

<strong>liquid</strong> <strong>crystal</strong> is sandwiched between two glass plates,<br />

forming a planar <strong>waveguide</strong> in the absence of an applied<br />

voltage. The anchoring conditions at the top and bottom<br />

glass surfaces determine the planar alignment of the molecules<br />

in the z-direction. To introduce a spatially periodic<br />

refractive modulation inside of NLC layer and to obtain a<br />

voltage-adjustable response, a set of regularly-spaced,<br />

transparent indium-tin-oxide (ITO) electrodes (as thin as<br />

50 nm) is placed on the top surface to apply a reorientational<br />

bias across the cell. The comb-shape of this electrode<br />

(with equal finger widths and spacing) allows us the<br />

formation of identical channels and provides confinement<br />

in the transverse x-y plane. In our samples, the electrode set<br />

had the periods ranging from 4 to 8 µm depending on the<br />

cell thickness.<br />

* e-mail: kasia@if.pw.edu.pl<br />

Opto-Electron. Rev., 13, no. 2, 2005 K.A. Brzd¹kiewicz 107


<strong>Nematic</strong> <strong>liquid</strong> <strong>crystal</strong> <strong>waveguide</strong> <strong>arrays</strong><br />

Fig. 1. Sketch of the <strong>liquid</strong> <strong>crystal</strong> <strong>waveguide</strong> array. The period of<br />

the electrode distribution L varies from 4 to 8 µm and the cell<br />

thickness d from 3 to 7 µm (depending on the cell).<br />

2. Numerical results<br />

The application of an appropriate voltage causes the reorientation<br />

of the NLC molecules which, as a result of free-<br />

-energy minimization, tend to align parallel to the direction<br />

of the electric field. The angular molecular distribution can<br />

be theoretically calculated from the Euler-Lagrange equation<br />

describing reorientation through an applied bias [21].<br />

The numerical results shown in this communication were<br />

obtained for the <strong>liquid</strong> <strong>crystal</strong> PCB, with low-frequency<br />

anisotropy De if = 115 . . Owing to a new arrangement of the<br />

orientation angle, a periodic spatial modulation of the refractive<br />

index occurs. Figure 2 shows an example of refractive<br />

index distribution (for x-polarized light) obtained in<br />

the 5-µm-thick sample and for electrode period of 6 µm.<br />

Figure 2(b) presents numerical results of refractive index<br />

distribution in the middle of the NLC layer (i.e., for x =0)<br />

and three different values of bias. As one can see, the refractive<br />

index increases mainly in the regions under the<br />

electrodes and thereby the channel <strong>waveguide</strong>s are well defined<br />

through lateral confinement. In general, the higher refractive<br />

index is obtained for a higher bias, and better light<br />

guiding is expected. In this case, the coupling length, defined<br />

as the propagation distance at which whole energy<br />

goes from one <strong>waveguide</strong> to the neighbouring ones [23],<br />

gets longer.<br />

Numerical simulations of light propagating in the NLC<br />

layer were carried out with BPM injecting a TM-polarized<br />

Gaussian beam. Additionally, in each propagation step the<br />

molecular orientation distribution was recalculated based<br />

on minimization of the total free-energy of the <strong>liquid</strong> <strong>crystal</strong><br />

under the influence of both the low-frequency (bias) and<br />

the optical electric field [21]. The refractive indices characteristic<br />

of PCB are n 0 = 1.52 and n e = 1.69 at a wavelength<br />

of 1064 nm. Figure 3 displays light propagation in the linear<br />

(low input optical power) regime. In this case, the electric<br />

bias plays a crucial role in the reorientation process,<br />

leading to the periodic modulation of a refractive index, as<br />

shown in Fig. 2. While standard diffraction (in the homogenous<br />

medium) takes place in the transverse y-z plane in the<br />

absence of bias, discrete diffraction [1,2,22] sets-in and can<br />

be tuned in strength as the voltage increases and the array<br />

is formed. As a consequence of discrete diffraction, energy<br />

redistributes among the guiding channels, as clearly visible<br />

in Figs. 3(b)–3(e) showing the cross-sections in x = 0 and<br />

beam transverse profiles in the x-y plane along propagation.<br />

In the analysed geometry, the optimisation of such parameters<br />

as the optical properties (especially birefringence)<br />

of NLC, the NLC layer thickness, the electrode spacing and<br />

width, allows us to obtain properly coupled and well defined<br />

<strong>waveguide</strong>s. Moreover, the magnitude of discrete diffraction<br />

can be easily tuned by controlling the voltage.<br />

Figure 4 shows how the coupling between channels and<br />

the resulting discrete diffraction changes with the applied<br />

voltage. An increase in bias reduces angular divergence of<br />

the beam; light is more confined within the core region of<br />

the channels, and the coupling length increases. As intuitive,<br />

by increasing the bias it is possible to progressively<br />

Fig. 2. Spatial distribution of the refractive index for light linearly polarized along x, NLC layer 5-µm thick and electrode width and spacing<br />

of 3 µm. Three-dimensional map for the bias V = 1.35 V (a), refractive index in the middle of the NLC cell (i.e., for x = 0) for three different<br />

applied voltages (b).<br />

108 Opto-Electron. Rev., 13, no. 2, 2005 © 2005 COSiW SEP, Warsaw


7 th International Workshop on Nonlinear Optics Applications<br />

Fig. 3. Numerical results for a low-power Gaussian optical input of waist 2 µm and wavelength 1064 nm, for a 5-µm-thick NLC-cell,<br />

electrode period of 6 µm and applied voltage V = 1.35 V. Discrete diffraction for the beam launched into one input <strong>waveguide</strong> (a). Beam<br />

transverse profiles (for x = 0) (b) and (c), spatial distribution of light intensity (d) and (e) after 1- and 2-mm propagation distances,<br />

respectively.<br />

reduce light transfer between channels. By using a proper<br />

bias the difference between refractive indices in each channel<br />

and in the gaps between the finger electrodes can be<br />

maximized. However, the saturating character of this reorientation<br />

process causes an excessive voltage to flatten the<br />

index profile and thereby weaken the spatial definition of<br />

each channel. The non local response of NLC, in fact, is<br />

such that an electric field increases the index not only under<br />

electrodes but also between them. For high enough applied<br />

voltages, the NLC elastic response mediates reorientation<br />

in the regions between neighbouring <strong>waveguide</strong>s,<br />

thereby reducing the refractive index modulation. This is<br />

also visible in Fig. 5(a) where, for a fixed cell thickness,<br />

the coupling length first increases with applied voltage until<br />

saturation and non locality makes it decrease again.<br />

Figure 5 shows how discrete diffraction can be modified<br />

by varying the sample geometry. By reducing the<br />

thickness of the NLC layer it is possible to obtain better defined<br />

<strong>waveguide</strong>s, i.e., an increased coupling length,<br />

Fig. 5(b). Figure 5(c) shows how the light beam injected in<br />

a 2.5-µm-thick sample for low bias (1.2 V) is confined<br />

within 3 channels after 1-mm propagation. Conversely, for<br />

thicker NLC layers, the change in coupling length versus<br />

voltage is negligible and discrete diffraction is hardly affected.<br />

As expected, the coupling length can be also reduced by<br />

decreasing the electrode periodicity. However, a close<br />

proximity of the stripe electrodes reduces the coupling<br />

length dependence with bias. For example, in a 5-µm-thick<br />

sample with 4-µm electrode period, by varying the voltage<br />

from 1.1 to 1.9 V, the change in coupling distance is<br />

around 60 µm. The analyses demonstrate that, in the process<br />

of sample designing, it is important to find the optimum<br />

ratio between electrode width and sample thickness<br />

in order to obtain a well tunable structure for suitable values<br />

of the applied voltage.<br />

Fig. 4. Linear coupling between channels and its dependence on applied voltage for NLC thickness of 5 µm and electrode period of 6 µm.<br />

The optical intensity distribution is colour-inverted (i.e., darker corresponds to a higher intensity) and evaluated after 1 mm propagation.<br />

Opto-Electron. Rev., 13, no. 2, 2005 K.A. Brzd¹kiewicz 109


<strong>Nematic</strong> <strong>liquid</strong> <strong>crystal</strong> <strong>waveguide</strong> <strong>arrays</strong><br />

Fig. 5. Linear coupling between channels versus NLC thickness. Dependence of a coupling length on applied voltage for three cell<br />

thicknesses d (a). Coupling length versus NLC thickness at a voltage V = 1.2 V (b). Spatial intensity distribution calculated after 1 mm<br />

propagation and V = 1.2 V for different values of NLC thickness d (c).<br />

Going from the linear to the nonlinear regime, as the<br />

power of the propagating beam increases, the refractive index<br />

of the excited <strong>waveguide</strong>s is modified by the nonlinearity,<br />

and discrete diffraction can be effectively counteracted<br />

by the power-induced detuning of the launch channel.<br />

As shown in Fig. 6, as the light power gets higher, the<br />

beam narrows in space. Finally, when the input beam is intense<br />

enough, nonlinearity completely balances diffraction<br />

and light propagates (transversely) localized in a limited region<br />

of the array or, for high enough excitations, in a single<br />

channel. Such a beam, which propagates maintaining an invariant<br />

transverse profile, is called discrete spatial soliton<br />

[1,2,18,22]. It is possible to consider either partially localized<br />

or “wide” solitons with energy in a few channels, or<br />

totally localized “narrow” solitons, with energy essentially<br />

confined in the input <strong>waveguide</strong>. It is worth to underline<br />

that, due to light confinement, discrete systems require less<br />

power for solitons than comparable slab <strong>waveguide</strong> or bulk<br />

media.<br />

Fig. 6. Beam distribution versus input beam power calculated for V<br />

= 1.35 V after 1-mm propagation. Discrete diffraction is limited<br />

and discrete spatial solitons are generated due to increase in input<br />

light power.<br />

3. Experimental results<br />

We performed experiments in NLC cell of a thickness of<br />

5–6 µm and with an array period of 6 µm. The beam from a<br />

Nd:YAG laser (l = 1064 nm) was focused to a diameter of<br />

approximately 4 µm at the entrance of the sample. The<br />

beam evolution in the NLC cell was observed through a<br />

CCD camera detecting the photons scattered above the cell.<br />

Photographs were taken at various biases, as shown in<br />

Fig. 7. The pictures show the change in discrete beam propagation<br />

for a fixed input power of 5 mW and a propagation<br />

distance of 1 mm. As one can see, by increasing the applied<br />

voltage the coupling length becomes longer and beam spatial<br />

localization is enhanced. For low voltages [Figs. 7(a)<br />

and 7(b)], the beam is guided in a few channels, whereas<br />

V = 1.45 V it essentially propagates in just one of them.<br />

As predicted theoretically, however, spatial light localization<br />

can be also obtained by increasing the beam power,<br />

leading to discrete soliton generation.<br />

Figure 8 shows the experimental photographs of light<br />

propagation depending on optical power in the structure.<br />

When no voltage is applied, the beam diffracts in the y-z<br />

plane, progressively reducing its intensity and becoming<br />

hardly visible after a few Rayleigh ranges, see Fig. 8(a).<br />

Figure 8(b) shows discrete diffraction of the excitation injected<br />

into a single channel for a bias V = 1.1 V, large<br />

enough to create an array of well-defined channels. In<br />

z = 1.4 mm, the input has spread over 13 adjacent <strong>waveguide</strong>s.<br />

Compared to standard diffraction in a homogeneous<br />

1D geometry [planar <strong>waveguide</strong>, Fig. 8(a)], discrete<br />

diffraction [Fig. 8(b)] gives rise to the smaller divergence.<br />

Figures 8(c)–8(e) present the successive changes of the<br />

110 Opto-Electron. Rev., 13, no. 2, 2005 © 2005 COSiW SEP, Warsaw


7 th International Workshop on Nonlinear Optics Applications<br />

4. Conclusions<br />

Fig. 7. Experimental results showing a 5-mW Nd:YAG beam after<br />

1-mm propagation for three different applied voltages.<br />

character of discrete propagation obtained due to increment<br />

of the beam evolution versus optical power. As expected,<br />

a high excitation mismatches the input <strong>waveguide</strong>, resulting<br />

in a “narrow” discrete soliton [Fig. 8(e)]. Figures<br />

8(f)–8(i) show the corresponding beam transverse profiles<br />

in z = 1.4 mm.<br />

In conclusions, we have presented nonlinear <strong>waveguide</strong><br />

<strong>arrays</strong> in undoped nematics. The investigated geometry<br />

takes advantage of both the electro-optic and all-optical<br />

response of <strong>liquid</strong> <strong>crystal</strong>s in a planar arrangement, offering<br />

considerable flexibility both in the linear and nonlinear<br />

regimes by voltage tuning the relevant parameters.<br />

The giant nonlinear response characteristic of molecular<br />

reorientation in such medium, allows for observation of<br />

discrete spatial solitons at mW levels, paving the way to<br />

multifunctional routers and signal processors in <strong>liquid</strong><br />

<strong>crystal</strong>s. The theoretical predictions are in excellent<br />

agreement with the experimental results, and further<br />

work is underway to enlighten a variety of additional fascinating<br />

phenomena including gap solitons and discrete<br />

breathers.<br />

Acknowledgments<br />

This work was partially supported by the State Committee<br />

for Scientific Research under grant No. 4 T11B 058 25.<br />

Fig. 8. Linear propagation (standard diffraction) of the beam in the homogenous planar <strong>waveguide</strong> (i.e., at zero voltage) (a). Light<br />

propagation (for voltage of 1.1 V) for different input powers of the beam launched into one channel (b–e). Acquired transverse profiles<br />

(solid line) in z = 1.4 mm, corresponding to the cases (b–e) (f–i). The dashed-line graphs are calculated profiles matching the experimental<br />

conditions.<br />

Opto-Electron. Rev., 13, no. 2, 2005 K.A. Brzd¹kiewicz 111


<strong>Nematic</strong> <strong>liquid</strong> <strong>crystal</strong> <strong>waveguide</strong> <strong>arrays</strong><br />

References<br />

1. F. Lederer, S. Darmanyan, and A. Kobyakov, “Discrete<br />

solitons”, in Spatial Solitons, edited by S. Trillo and<br />

W. Torruellas, Wiley, New York, 2002.<br />

2. A.A. Sukhorukov, Y.S. Kivshar, H.S. Eisenberg, and Y.<br />

Silberberg, “Spatial optical solitons in <strong>waveguide</strong> <strong>arrays</strong>”,<br />

IEEE J. Quantum Electron. 39, 31 (2003).<br />

3. D.N. Christodulides and E.D. Eugenieva, “Blocking and<br />

routing discrete solitons in two-dimensional networks of<br />

nonlinear <strong>waveguide</strong> <strong>arrays</strong>”, Phys. Rev. Lett. 87, 233901<br />

(2001); E.D. Eugenieva, N.K. Efremidis, and D.N. Christodoulides,<br />

“Minimizing bending losses in two-dimensional<br />

discrete soliton networks”, Opt. Lett. 26, 1876 (2001), “Design<br />

of switching junctions for two-dimensional discrete<br />

soliton networks”, Opt. Lett. 26, 1978 (2001), “Minimizing<br />

bending losses in two-dimensional discrete soliton networks:<br />

errata”, Opt. Lett. 27, 369 (2002).<br />

4. R. Morandotti, U. Peschel, J.S. Aitchison, H.S. Eisenberg,<br />

and Y. Silberberg, “Dynamics of discrete solitons in optical<br />

<strong>waveguide</strong> <strong>arrays</strong>”, Phys. Rev. Lett. 81, 2726 (1999).<br />

5. W. Królikowski and Y.S. Kivshar, “Soliton-based optical<br />

switching in <strong>waveguide</strong> <strong>arrays</strong>”, J. Opt. Soc. Am. B13, 876<br />

(1996).<br />

6. O. Bang and P. Miller, ”Exploiting discreteness for switching<br />

in <strong>waveguide</strong> <strong>arrays</strong>”, Opt. Lett. 21, 1105 (1996).<br />

7. D.N. Christodoulides and R.I. Joseph, “Discrete self-focusing<br />

in nonlinear <strong>arrays</strong> of coupled <strong>waveguide</strong>s”, Opt. Lett.<br />

13, 794 (1988).<br />

8. C. Etrich, F. Lederer, B. Malomed, T. Peschel, and U.<br />

Peschel, “Optical solitons in media with a quadratic<br />

nonlinearity”, in Progress in Optics, Vol. 41, pp. 483, edited<br />

by E. Wolf, Elsevier Science Publishers 2000; T. Peschel,<br />

U. Peschel, and F. Lederer, “Discrete bright solitary waves<br />

in quadratically nonlinear media”, Phys. Rev. E57, 1127<br />

(1998).<br />

9. J. Meier, G.I. Stegeman, Y. Silberberg, R. Morandotti, and<br />

J.S. Aitchison, “Nonlinear optical beam interactions in<br />

<strong>waveguide</strong> <strong>arrays</strong>”, Phys. Rev. Lett. 93, 093903 (2004).<br />

10. R. Morandotti, H.S. Eisenberg, Y. Silberberg, M. Soler, and<br />

J.S. Aitchison, “Self-focusing and defocusing in <strong>waveguide</strong><br />

<strong>arrays</strong>”, Phys. Rev. Lett. 86, 3296 (2001).<br />

11. H.S. Eisenberg, Y. Silberberg, R. Morandotti, A.R. Boyd,<br />

and J.S. Aitchison, “Discrete spatial optical solitons in<br />

<strong>waveguide</strong> <strong>arrays</strong>”, Phys. Rev. Lett. 81, 3383 (1998); “Diffraction<br />

management”, Phys. Rev. Lett. 85, 1863 (2000).<br />

12. P. Millar, J.S. Aitchison, J.U. Kang, G.I. Stegeman, A.<br />

Villeneuve, G.T. Kennedy, and W. Sibbett, J. Opt. Soc. Am.<br />

B14, 3224 (1997).<br />

13. D. Cheskis, S. Bar-Ad, R. Morandotti, J.S. Aitchison, H.S.<br />

Eisenberg, Y. Silberberg, and D. Ross, “Strong spatiotemporal<br />

localization in a silica nonlinear <strong>waveguide</strong> array”,<br />

Phys. Rev. Lett. 91, 223901 (2003).<br />

14. T. Pertsch, U. Peschel, F. Lederer, J. Burghoff, M. Will, S.<br />

Nolte, and A. Tunnermann, “Discrete diffraction in two-dimensional<br />

<strong>arrays</strong> of coupled <strong>waveguide</strong>s in silica”, Opt.<br />

Lett. 29, 468 (2004).<br />

15. N.K. Efremidis, S. Sears, D.N. Christodoulides, J.W.<br />

Fleischer, and M. Segev, “Discrete solitons in photorefractive<br />

optically induced photonic lattices”, Phys. Rev. E66,<br />

046602 (2002).<br />

16. J.W. Fleisher, T. Carmon, M. Segev, N.K. Efremidis, and<br />

D.N. Christodoulides, “Observation of discrete solitons in<br />

optically induced real time <strong>waveguide</strong> <strong>arrays</strong>”, Phys. Rev.<br />

Lett. 90, 023902 (2003); J.W. Fleisher, M. Segev, N.K.<br />

Efremidis, and D.N. Christodoulides, “Observation of<br />

two-dimensional discrete solitons in optically induced nonlinear<br />

photonic lattices”, Nature 422, 147 (2003).<br />

17. R. Iwanow, R. Schiek, G.I. Stegeman, T. Pertsch, F.<br />

Lederer, Y. Min, and W. Sohler, “Observation of discrete<br />

quadratic solitons”, Phys. Rev. Lett. 93, 113902 (2004).<br />

18. A. Fratalocchi, G. Assanto, K.A. Brzd¹kiewicz, and M.A.<br />

Karpierz, “Discrete propagation and spatial solitons in nematic<br />

<strong>liquid</strong> <strong>crystal</strong>s”, Opt. Lett. 29, 1530 (2004).<br />

19. I.C. Khoo and S.T. Wu, Optics and Nonlinear Optics of Liquid<br />

Crystals, World Scientific Publ., Singapore 1997; F.<br />

Simoni, Nonlinear Optical Properties of Liquid Crystals,<br />

World Scientific Publ., London, 1997.<br />

20. M. Peccianti, G. Assanto, A. De Luca, C. Umeton, and I.C.<br />

Khoo, “Electrically assisted self-confinement and waveguiding<br />

in planar nematic <strong>liquid</strong> <strong>crystal</strong> cells”, Appl. Phys.<br />

Lett. 77, 7 (2000); M. Peccianti, K.A. Brzd¹kiewicz, and G.<br />

Assanto, “Nonlocal spatial soliton interactions in nematic<br />

<strong>liquid</strong> <strong>crystal</strong>s”, Opt. Lett. 27, 1460 (2002); G. Assanto and<br />

M. Peccianti, “Spatial solitons in nematic <strong>liquid</strong> <strong>crystal</strong>s”,<br />

IEEE J. Quantum Electron. 39, 13 (2003); M.A. Karpierz,<br />

“Solitary waves in <strong>liquid</strong> <strong>crystal</strong>line <strong>waveguide</strong>s”, Phys.<br />

Rev. E66, 036603 (2002); M.A. Karpierz, M. Sierakowski,<br />

M. Œwi³³o, and T.R. Woliñski, “Self focusing in <strong>liquid</strong> <strong>crystal</strong>line<br />

<strong>waveguide</strong>s”, Mol. Cryst. Liq. Cryst. 320, 157<br />

(1998).<br />

21. N.V. Tabiryan, A.V. Sukhov, and B.Ya. Zel'dovich,<br />

“Orientational optical nonlinearity of <strong>liquid</strong> <strong>crystal</strong>s”, Mol.<br />

Cryst. Liq. Cryst. 136, 1 (1986); P.J. Collings, M. Hird, Introductions<br />

to Liquid Crystals – Chemistry and Physics,<br />

Taylor & Francis, London 1997.<br />

22. F. Lederer and Y. Silberberg, “Discrete solitons”, Opt. Photon.<br />

News 2, 49 (2002); D.N. Christodoulides, F. Lederer,<br />

and Y. Silberberg, “Discretising light behaviour in linear<br />

and nonlinear <strong>waveguide</strong> lattices”, Nature 424, 817 (2003).<br />

23. A. Yariv, Optical Electronics in Modern Communications.<br />

Oxford Press, New York 1997.<br />

112 Opto-Electron. Rev., 13, no. 2, 2005 © 2005 COSiW SEP, Warsaw

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!